Int. J. Mol. Sci. 2014, 15, 6974-6989; doi:10.3390/ijms15046974 OPEN ACCESS

International Journal of

Molecular Sciences ISSN 1422-0067 www.mdpi.com/journal/ijms Article

Synthesis of Environmentally Friendly Highly Dispersed Magnetite Nanoparticles Based on Rosin Cationic Surfactants as Thin Film Coatings of Steel Ayman M. Atta 1,2,†, Gamal A. El-Mahdy 1,3,†,*, Hamad A. Al-Lohedan 1,† and Sami A. Al-Hussain 1,† 1

2 3



Surfactant Research Chair, Chemistry Department, College of Science, King Saud University, P.O. Box-2455, Riyadh 11451, Saudi Arabia; E-Mails: [email protected] (A.M.A.); [email protected] (H.A.A.-L.); [email protected] (S.A.A.-H.) Petroleum Application Department, Egyptian Petroleum Research Institute, Cairo 11795, Egypt Chemistry Department, Faculty of Science, Helwan University, Helwan, Cairo 11727, Egypt These authors contributed equally to this work.

* Author to whom correspondence should be addressed; E-Mail: [email protected]; Tel.: +966-11-467-5998; Fax: +966-11-467-5992. Received: 3 February 2014; in revised form: 5 March 2014 / Accepted: 4 April 2014 / Published: 22 April 2014

Abstract: This work presents a new method to prepare monodisperse magnetite nanoparticles capping with new cationic surfactants based on rosin. Core/shell type magnetite nanoparticles were synthesized using bis-N-(3-levopimaric maleic acid adduct-2-hydroxy) propyl-triethyl ammonium chloride (LPMQA) as capping agent. Fourier transform infrared spectroscopy (FTIR) was employed to characterize the nanoparticles chemical structure. Transmittance electron microscopies (TEM) and X-ray powder diffraction (XRD) were used to examine the morphology of the modified magnetite nanoparticles. The magnetite dispersed aqueous acid solution was evaluated as an effective anticorrosion behavior of a hydrophobic surface on steel. The inhibition effect of magnetite nanoparticles on steel corrosion in 1 M HCl solution was investigated using potentiodynamic polarization curves and electrochemical impedance spectroscopy (EIS). Results obtained from both potentiodynamic polarisation and EIS measurements reveal that the magnetite nanoparticle is an effective inhibitor for the corrosion of steel in 1.0 M HCl solution. Polarization data show that magnetite nanoparticles behave as a mixed type inhibitor. The inhibition efficiencies obtained from potentiodynamic polarization and EIS methods are in good agreement.

Int. J. Mol. Sci. 2014, 15

6975

Keywords: magnetite; nanoparticles; cationic surfactant; rosin; encapsulation; EIS; polarization

1. Introduction Synthesis of new bioactive materials from renewable natural resources has become a rapidly growing area to replace or partially replace environmentally and energy unfavorable toxic materials derived from petrochemicals. Rosin, a robust inexhaustible raw material from pine tree is mixture of compounds in which 90% is acidic called resin acids with the formula of C19H29COOH and the last 10% is neutral compounds. Resin acids have characteristic bulky hydrophenanetherene and two activity masses-a double-bond conjugate system and a carboxyl group that tend to participate readily in reactions that provides resin acids [1,2]. Dehydrogenated rosin and hydrogenated rosin are bioactive materials and widely used in many fields such as medicine [3,4] and antifungal activity [5]. It is well published that quaternary ammonium had robust and broad spectrum anti-fungal activity [6]. A series of Diels-Alder adducts of resin acid with acrylic acid were synthesized which had great antimicrobial activity against different kinds of bacteria by filter paper method [7,8]. As an inexpensive natural product, fungicides related to rosin have superiority. Moreover, highly active surface modified cheap rosin materials can be advantageous in the production of new materials. This is attributed to its cheap price and low toxicity when compared to the petroleum-based materials. We started to modify rosin chemical structure which can be converted to highly surface active materials to apply as additives for petroleum crude oil applications [9–18]. The corrosion protection of metals and steel has recently gained momentum using nanotechnology [19–22]. A polymer Nano composite coating can effectively protect the underlying substrate by combining the benefits of organic polymers such as water resistance and elasticity to that of advanced inorganic materials such as hardness [20]. Nanotechnology has been employed to reduce the impact of corrosive environments through the alternation of the steel/electrolyte interface (e.g., formation of Nano-composite thin film coatings on steel). Magnetite is one of the most promising nanoparticles employed in anticorrosion coatings [23,24]. Recently, magnetite (Fe3O4) nano- and micro-particles attracted more attention in material chemistry, separation technology, biology, and biomedicine, leading to a number of potential applications [25–29]. It is chemically stable, nontoxic and noncarcinogenic [30]. Many synthetic routes have been reported in the literature for chemical synthesis of these particles such as the sol-gel [31], microemulsion [32], sonochemical [33], ultrasonic spray pyrolysis [34], and microwave plasma [35]. Each preparation method has its advantages and disadvantages, which primarily relate to particles’ size distribution, production scale, and cost. Two important synthesis routes are the thermal decomposition [36] and the chemical co-precipitation reduction [37] methods. A characteristic and serious problem is in assembling and stabilizing magnetic particles as it has high affinity to agglomerate, which is an obstacle to its application and activity. This problem can be resolved if the particles are dispersed using proper dispersing agent. The surface-modification techniques that can transform these materials into valuable finished products are an important part of nanomaterials. Moreover, they allow for the manipulation of surface functions to meet diverse application requirements. Surface modification of nanomaterials with functional polymers is an important research

Int. J. Mol. Sci. 2014, 15

6976

area in polymer science owing to the wide applications. Here, we attempted to solve the synthesis of stabilized and dispersed Fe3O4 nanoparticles using chemical reduction method at ambient temperature using new cationic surfactants based on rosin as dispersing agent. Modified magnetite particles have advantages over normal organic inhibitors due to the formation of uniform thin films on the surface of steel, which entirely cover the entire surface without any defects. In the present work, it was thought valuable to investigate the application of magnetite coated with new cationic surfactants as a corrosion inhibitor for steel in an aggressive acidic medium (1 M HCl) using potentiodynamic polarization curves and electrochemical impedance spectroscopy (EIS). 2. Results and Discussion 2.1. Preparation of Cationic Surfactant The synthetic route of LPMQA cationic surfactant was illustrated in Scheme 1. In this scheme the rosin acids was isomerized to leveopimaric acid (LPA) by heating rosin according the reported route as explained in the experimental section. The Diels Alder adduct between levopimaric acid with MA (LPMA) was reacted with oxolylchloride to produce leveopimaric chloride maleic anhydride adduct. The produced acid chloride reacted with EC in the presence of acid acceptor based on triethylamie to produce 2-hydroxy-3-leveopimaric maleic anhydride propyl chloride ester (LPME). The produced chloride ester of LLPME was reacted with TEM to produce bis-N-(3-levopimaric maleic acid adduct-2-hydroxy) propyl-triethyl ammonium chloride (LPMQA) as cationic surfactant. Scheme 1. Synthetic routs of LPMQA. COOH

COOH

O

COOH

O

MA

heating

O

glacial acetic LPMA

LPA

Oxalyl chloride

AbA OH COOCH2CHCH2Cl

COCl

O

COOH

CH2Cl

O O

COOH

O

LPME LPCMA

N(CH2CH3)3 OH

Cl

+

CH2CH3

COOCH2CHCH2N COOH

CH2CH3 CH2CH3

COOH LPMQA

IR spectra of compounds of, LPME and LPMQA are shown in Figure 1. Compared with the raw material AbA, the absorbance at 3430 cm−1 in the other three spectra heightens as the three-membered ring of epichlorohydrin opened and –OH is produced. The absorbance at 1700 cm−1 of C=O stretching vibration in spectrum of AbA converts to 1720 cm−1 of LPME and LPMQA after the esterification, and

Int. J. Mol. Sci. 2014, 15

6977

quaternization. The appearance of new peak around 1371 cm−1 was assigned to C–N stretching vibration. Due to the quaternization (Figure 1c), a small intense peak at 2365.72 cm−1 has appeared, which is the characteristic absorption peak of quaternary ammonium groups. The chemical structure of LPMQA surfactant was also confirmed by 1H NMR analysis. The spectrum of LPMQA surfactant was represented in Figure 2. The chemical shifts (δ) are 5.39 (s, 1H; OH), 4.71–4.33 (m, 2H, –CH–OH), 4.25–4.08 (m, 4H, –COO–CH2), 3.99–3.85 (m, 2H, –CH–OH), 3.59–3.30 (m, 16H, –CH2–N–), 2.50–2.37 (m, 2H), 2.22–2.10 (m, 2H), 1.95–1.39 (m, 17H), 1.38–1.24 (m, 18H, –N–CH3), 1.21–1.01 (m, 10H). Figure 1. FTIR spectra of: (a) rosin AbA; (b) LPME; (c) LPMQA.

Figure 2. 1H NMR spectrum of LPMQA.

2.2. Surface Modification of Magnetite Surface modification of nanoparticle was used to prevent agglomeration during particle synthesis by forming a dense protecting monolayer on the particle to provide effective stabilization. In the present work, we selected new cationic surfactants to stabilize the magnetite nanoparticles. The magnetite was prepared previously [38,39] at room temperature by oxidation reduction reaction of ferric chloride in the presence of potassium iodide at pH 9 using ammonia or sodium hydroxide aqueous solution and

Int. J. Mol. Sci. 2014, 15

6978

stabilized with nonionic surfactant. In the present work, the preparation and stabilization of magnetite using cationic surfactant is one of the research objectives. A schematic probable representation of the adsorption mechanism of LPMQA cationic surfactants is shown in Scheme 2. In this respect, assembling a monolayer of the LPMQA cationic surfactant molecules on the nanoparticle surface produces oil-based magnetic fluid (Scheme 2a). The aqueous-based magnetic fluid is produced by assemblies of bilayers of LPMQA surfactant molecules (Scheme 2b). The adsorption of the LPMQA cationic surfactants on the surface of Fe3O4 NPs are driven by electrostatic interactions (Scheme 2c). The charge-stabilized mechanism is most favorable to explain the interaction between magnetite surface and LPMQA cationic surfactants. The presence of LPMQA cationic surfactant, its molecules interact with the negatively charged surface of Fe3O4 NPs and create a positive surface at a pH = 9.0 via coating of the surface of Fe3O4 NPs. Therefore, at alkaline pH = 9 is found with increasing amounts of LPMQA cationic surfactant, the adsorption efficiency of the surfactant will increase on the Fe3O4 NPs surfaces. It was previously reported that [40] cetyltrimethylammonium bromide coated magnetite NPs (CTAB-coated Fe3O4 NPs) cannot interact with the positively charged Fe3O4 NPs due to electrostatic repulsion under acidic conditions. Moreover, at high concentrations of CTAB, the adsorption efficiency decreased due to the formation of CTAB (micelles and/or) ion-pairing between excess amounts of CTAB which reduce the interaction with positively charged surface of Fe3O4 NPs. Accordingly, If concentration of cationic surfactant exceeds the critical micelle concentration, micelle formation will occur. This will reduce the amount of LPMQA cationic surfactant available for particles stabilization. Attempts were made to prevent micelle formation by adding LPMQA cationic surfactant in portions. Stabilization of magnetite nanoparticles with LPMQA cationic surfactant at high pH is attributed to adsorption of LPMQA cationic ions at the particle surface and subsequent charge stabilization of the magnetic particles. Scheme 2. Schematic illustration of the adsorption mechanism of (a) monolayer assembly; (b) double layer assembly and (c) chemical interaction of LPMQA onto Fe3O4 NPs.

(a)

(b)

(c)

Int. J. Mol. Sci. 2014, 15

6979

The IR spectra of pure Fe3O4 NPs and LPMQA-coated Fe3O4 NPs are shown in Figure 3. The band at 3455 cm−1 is attributed to the stretching vibrations of –OH, which is assigned to surface OH groups of Fe3O4 NPs (Figure 3a). The band at 571 cm−1 is attributed to the Fe–O band vibration of Fe3O4. Also, the band at 1380 cm−1 (Figure 3b) is attributed to C–N band and the peaks at 2885 and 2925 cm−1 are attributed to two different C–H bands vibration of LPMQA. The IR spectra show that Fe3O4 NPs surface was well modified by LPMQA. The TEM micrographs (Figure 4) of the Fe3O4− LPMQA coated NPs demonstrate the high quality of the as-synthesized nanoparticles. The Fe3O4− LPMQA coated NPs analyzed by TEM showed a spherical shape with a narrow size distribution (Figure 4a,b). In order to further understand the mechanism of the surfactants adsorption on the bare magnetite surface we measured the adsorption isotherm of LPMQA rosin surfactants on the hydrophilic magnetite particles in water as illustrated in the experimental section. The absorbance values of the absorbed LPMQA on the magnetite were determined vs. concentrations and plotted to fit the equation Y = a + bX; where Y represents the absorbance and X is the concentration of the surfactants. The equations (with r2 = 0.9997) were obtained as absorbance = 0.001 + 0.163 C. The adsorption activity of LPMQA on the surface of hydrophilic magnetite was measured and listed in Table 1. The data show that the adsorbed amounts at LPMQA displays two plateaus. The first at low concentrations corresponds to the interaction between cationic nitrogen groups and hydroxyl groups of the magnetite as illustrated in Figure 3. It is thus suggested that the initial adsorption of LPMQA molecules on the particles is driven by electrostatic interaction. Increasing the concentration of LPMQA increases monotonously just before the second plateau is reached at 1.9 mM which was determined as the optimum concentration of LPMQA surfactant to adsorb onto the dispersed bare magnetite nanoparticles (concentration 10%). It is suggested that the LPMQA molecules form a monolayer on the particle surface with their hydrocarbon chains towards water, i.e., the adsorption is via the interaction of the LPMQA nitrogen head groups with the OH of magnetite surface. These results indicated that the LPMQA rosin surfactant molecules first adsorb as individual molecules through interaction with OH groups of magnetite, and then associate into hemimicelles. In the hemimicelles, the surfactants attach their head-groups to the magnetite surface and orient their tail groups toward solution, thus forming hydrophobic patches on the surface [41]. Further adsorption then brings an increasing number of surfactants to the interface to form a complete second layer of bilayer on magnetite surfaces [41]. Figure 3. FTIR of spectra of: (a) pure Fe3O4 NPs; and (b) Fe3O4 NPs coated with LPMQA.

Int. J. Mol. Sci. 2014, 15

6980 Figure 3. Cont.

Figure 4. TEM micrograph (a) and histogram (b) of the as-prepared magnetite nanocrystals. The y-axis of the histogram represents the number of particles.

Table 1. Adsorption of the LPMQA prepared surfactants on the surface of the dispersed magnetite particle (10 wt %). Surfactant Concentrations wt % mol/L 0.288 0.0048 2.88 0.048 5.76 0.096

Adsorption Activity (%) 36.4 38.3 48.3

The image reveals the formation of monodisperse nanoparticles with an average size of 12.5 nm confirming that the proposed scheme is an effective method for the preparation of monodisperse magnetic nanoparticles. The magnetite nanoparticles appear as spherical dots with very narrow size distributions, typically 5%–10%. The spontaneously formed ordered assembly pattern is an additional confirmation of the monodispersity of the nanoparticles. The thickness of the shells is 3.7 nm. The thickness of the surfactant shell estimated using the difference between the magnetite core and the thickness of coated magnetite core shell as illustrated in the high resolution TEM micrograph at the bottom of Figure 4a. Figure 5 shows the XRD pattern of the synthesized NPs, which is quite identical to pure magnetite (JCPDS No. 07-0322), indicating that the sample has a cubic crystal system. Also, it can be seen that no

Int. J. Mol. Sci. 2014, 15

6981

characteristic peaks of impurities were observed which demonstrates that the designed reaction procedure is effective to synthesize pure Fe3O4 NPs. The calculated d-spacing, matches well to the major d-spacing of magnetite (Fe3O4) [42,43]. The average size of the magnetite nanoparticles was estimated as 11.7 nm using the Scherrer equation: d = 0.9 λ/Δ(2θ)cosθ, where d is the crystalline domain size, Δ(2θ) is the width at half-maximum of the strongest peak (311),and λ is the X-ray wavelength. The Fe3O4 NPs size is in good agreement with the average diameter measured from TEM (12.5 nm). Figure 5. XRD pattern of the synthesized (a) bare Fe3O4 NPs; and (b) Fe3O4coated with LPMQA.

(a)

(b)

2.3. Potentiodynamic Polarization The potentiodynamic polarization curves for corrosion of steel in 1 M HCl, solutions without and with magnetic nanparticles were shown in Figure 6. The presence of magnetic nanparticles causes a prominent shift for both anodic and cathodic curves to lower values of current densities. The anodic and cathodic reactions of steel are clearly inhibited with increasing the concentration of magnetic nanparticles. This result suggests that addition of magnetic nanparticles in 1 M HCl solutions significantly reduces anodic dissolution and retards the cathodic reaction of steel. The results clearly indicate that magnetic nanparticles act essentially as a mixed type inhibitor in 1 M HCl solutions. The inhibition action is caused by geometric blocking effect by a reduction of the active sites area on the surface of the corroding steel [44]. It can be concluded that both cathodic and anodic reactions of steel corrosion are drastically retarded by the addition of magnetic nanparticles in 1 M HCl solutions. Accordingly, the corrosion current density values are estimated accurately by extrapolating the cathodic and anodic linear region to the corrosion potential. The electrochemical corrosion parameters including corrosion current densities (icorr), corrosion potential (Ecorr), anododic Tafel slope (Ba), cathodic Tafel slope (Bc) and corresponding inhibition efficiency (IE%) are given in Table 2. The inhibition efficiency (IE%) was calculated from potentiodynamic polarization using the following relationship [45–47]: IE% = 1 − icorr (inh)/icorr (uninh) × 100

(1)

where icorr (uninh) and icorr (inh) are corrosion current density values in the absence and presence of inhibitor, respectively. The inhibition efficiencies (IE%) were also listed in Table 2. It is evident that IE%

Int. J. Mol. Sci. 2014, 15

6982

increases with the inhibitor concentration, due to an increase in the blocked area of active sites of the steel surface. The decrease of corrosion rate may be explained by the inhibitory action of inhibitor on cathodic and anodic branches of the polarization curves by blocking the anodic and cathodic active sites on the steel surface. Figure 6. Potentiodynamic polarisation curves for steel in 1.0 M HCl without and with different concentrations of inhibitor.

Table 2. Inhibition efficiency values for steel in 1 M HCl with different concentrations of inhibitor calculated by Polarization and Electrochemical Impedance Spectroscopy (EIS) methods at 25 °C.

Blank 50 ppm 150 250

Ba (mV) 147.00

Polarization Method Bc (mV) Ecorr (V) icorr (µA/cm2) 141.00 −0.4034 745

IE% -

EIS Method Rct (Ohm) Cdl (µF/cm2) 1.80 334

IE% -

93.63

128.85

−0.4050

191.17

74.34

6.90

189

73.91

112.70 113.53

147.85 113.00

−0.4156 −0.4496

167.8 89.40

77.40 88.00

8.07 15.01

152 136

77.89 88.00

2.4. Electrochemical Impedance Spectroscopy (EIS) Figure 7 shows the typical Nyquist diagram for corrosion of steel in 1 M HCl solution with and without inhibitor. The impedance diagrams consist of single semicircle capacitive loop indicating a charge-transfer process mainly controlling the corrosion of steel. The EIS plots can be interpreted by an equivalent circuit with a solution resistance (Rs), a charge transfer resistance (Rct) and a double layer capacitance (Cdl) as shown in Figure 8. The circuit employed allows the identification of both charge transfer resistance (Rct) and double layer capacitance (Cdl). The fit parameters were analyzed using Zview2 software and listed in Table 1. The inhibition efficiencies (IE%) for the corrosion of steel were calculated and also listed in Table 1 by means of the following equation [45–47]: IE% = 1 − R°ct/Rct(inh)

(2)

where R°ct and Rct(inh) are the values of the charge transfer resistance without and with inhibitor, respectively. As can be seen from the data presented in Table 2, the inhibition efficiencies increase

Int. J. Mol. Sci. 2014, 15

6983

significantly with the concentration of inhibitor. It can be concluded that magnetic nanoparticles exhibit effective inhibition for the corrosion of steel in 1 M HCl solution. As the inhibitor concentration increased, the Rct values increased, but the Cdl values tended to decrease. The decrease in Cdl value may be accounted to the adsorption of inhibitor on steel surface, which led to an increase in IE% [48–50]. In addition, the decrease in the Cdl values can be attributed to a decrease in local dielectric constant and/or an increase in the thickness of the electrical double layer, which enhance the adsorption of inhibitor on the steel surface. The inhibition efficiencies, calculated from potentiodynamic polarization curve and EIS measurements are in reasonably good agreement. The results of EIS and polarization measurements are in good agreement with that reported previously for improving the anticorrosive performance of steel with the addition of nanoparticles to corrosive medium [51–53]. Figure 7. Nyquist plot for steel in 1.0 M HCl without and with different concentrations of inhibitor.

Figure 8. Equivalent circuit used to fit EIS data.

3. Experimental Section 3.1. Materials Gum rosin was obtained from Wuzhou Chemical, China and used as received. Rosin acid (AbA) with acid number 165 mg KOH·g−1 was obtained from a commercial resin by crystallization from cooled concentrated acetone solutions and purified by recrystallization from the same solvent. The separation of the resin acids from rosin was carried out to increase the yield and to remove terpens. Maleic anhydride (MA), epichlorohydrin (EC), tetrahydrofuran (THF), acetate acid, triethylamie (TEA) and oxalyl

Int. J. Mol. Sci. 2014, 15

6984

chloride (C2Cl2O2) were purchased from Aldrich Chemical Co. Anhydrous ferric chloride (St. Louis, MO, USA) potassium iodide and aqueous (25%) ammonium hydroxide solution used as reagents for preparation of magnetite were purchased from Aldrich Chemical Co. (St. Louis, MO, USA). Leveopimaric maleic anhydride adduct (LPMA) was prepared as reported in the previous work [18]. 3.2. Preparation Methods 3.2.1. Synthesis of Bis-N-(3-levopimaric maleic acid adduct-2-hydroxy) Propyl-Triethyl Ammonium Chloride (LPMQA) LPMA (10 g, 0.025 mol) was dissolved in 50 mL THF and added to a round bottom flask. Oxalyl chloride (2.95 mL, 0.03 mol) was injected and stirred at 50 °C for 4 h. The gas generated in the reaction was absorbed by base solution. After removed of the unreacted oxalyl chloride by vacuum distillation, the LPMA chloride (LPCMA) was obtained. Then, epichlorohydrin (18.5 g, 0.2 mol) and TEA (0.32 g, 0.001 mol) were added dropwise and the reaction temperature was heated to 50 °C and kept for 4 h. A high viscosity liquid was obtained after reduced pressure distillation was used to remove the excess epichlorohydrin. Then the mixture was extracted by ether and water three times. Aqueous phase remained and distilled under reduced pressure for recrystallization from alcohol and acetone. The produced bis-N-(3-levopimaric maleic acid adduct-2-hydroxy) propyl-triethyl ammonium chloride (LPMQA) was obtained as light yellow powder has yield: 72%. 3.2.2. Preparation of Modified Magnetite Nanoparticles The modified magnetite nanoparticles were made by coating magnetite with LPMQA. In this respect, a solution of anhydrous FeCl3 (40 g) in distilled water (300 mL) was prepared and named aqueous solution A. Further, potassium iodide (13.2 g, 0.08 mol) was dissolved in distilled water (50 mL) to prepare an aqueous solution B. The aqueous solutions A and B are then mixed together at room temperature, stirred and allowed to reach equilibrium for one hour while bubbling with pure N2 to keep the mixture oxygen free throughout the preparation procedure. LPMQA (3 g) diluted in water (100 mL) was added after heating the reaction mixture at 40 °C under vigorous stirring, then NH4OH (25%, 200 mL) was added dropwise to the above solution. After that, the reaction mixture was stirred for another 0.5 h at 50 °C. Finally, the LPMQA-stabilized Fe3O4 nanoparticles were separated from the mixture by ultracentrifugation at 12,000 rpm, washed with water three times and with ethanol two times. 3.3. Characterization Fourier Transform infrared (FTIR) spectroscopic analysis of the samples was performed using a Spectrum One FTIR spectrometer (Perkin-Bhaskar-Elmer Co., Boston, MA, USA). 1 H NMR (400 MHz Bruker Avance DRX-400 spectrometer; (Bruker Analytische Messtechnik, Karlsruhe, Germany) was used to investigate the chemical structures of the prepared rosin derivatives using CDCl3 as solvent and tetramethylsilane (TMS) as an internal reference. The morphology and structure of the prepared magnetic nanoparticles were determined using high-resolution transmission electron microscopy (HR-TEM). HR-TEM images of the nanocomposites were recorded using a JEM-2100 F (JEOL, Tokyo, Japan) at an acceleration voltage of 150 kV. The particle size distributions

Int. J. Mol. Sci. 2014, 15

6985

(PSDs) were determined from TEM microphotographs on representative samples of more than 1000 particles of magnetic latexes and analyzed using the Bolero software (AQ Systems, Tyresö, Sweden). A magnetite nanoparticle solution (1 mg) was diluted into 10 mL of ethanol, sonicated for 5 min and the resulting magnetite dispersed solutions solution. A few drops of the magnetite dispersed in ethanol solution were placed onto a carbon coated copper grid and allowed to evaporate. XRD patterns of the samples were obtained using a diffractometer (MiniFlex X-ray diffractometer/PW 3710, Rigaku, Japan) equipped with Cu–Ka (λ = 0.154 nm) radiation. 3.4. Materials The experiments were performed with steel specimens with the following chemical composition (wt %): 0.14% C, 0.57% Mn, 0.21% P, 0.15% S, 0.37% Si, 0.06% V, 0.03% Ni, 0.03% Cr and the remainder Fe. Prior to experiments, the specimens were successively polished using emery papers from 100 to 4000 grade. Then, the specimens were washed with distilled water and degreased with acetone and dried with a cold air. 3.5. Measurement of Adsorption Isotherm of LPMQA on Magnetite Nanoparticles Bare magnetite particles (0.3 g; 2 wt %) were dispersed in 15 mL of double distilled water containing different concentrations of the prepared surfactants (ranged from 0.1–5 wt %) using an ultrasound probe for 2 min at 50 W output and the dispersions were put in a thermostatic chamber at 25 °C for 10 min. The suspended particles were separated by centrifugation (8000 rpm for 10 min), the concentration of the prepared surfactants in the supernatant was measured using a spectrophotometric method. This could be achieved by monitoring the intensity of the absorption at 285 nm. A calibration curve (based on Beer’s low) represents the variation of rosin surfactant absorbance vs. their concentrations. The absorbance values of the rosin blank solutions in the absence of silica were determined as blank absorbance. The concentration of the absorbed surfactants on magnetite particles was determined from absorbance differences between blank solution and surfactant concentrations in the supernatant (after separation of silica by centrifuge). The adsorption activity values can be calculated as; activity = [(Co − C) × 100/(Co)] where; C and Co are concentrations of rosin surfactants (mmol/L) in the presence and absence of magnetite particles, respectively. 3.6. Electrochemical Measurements A three-electrode system including a steel as working electrode (WE), Pt sheet as counter electrode (CE) and saturated calomel electrode (SCE) as a reference electrode (RE) were used for electrochemical measurement. Potentiodynamics polarization measurements were conducted with a computer-controlled (potentiostat/galvanostat) Solartron 1470E system (Solartron Analytical, Cambridge, UK). Impedance measurements were carried out using Solartron 1470E system (potentiostat/galvanostat) with Solartron 1455A (Solartron Analytical, Cambridge, UK) as frequency response analyzer in the frequency range of 10 mHz–100 kHz using a 10 mV peak-to-peak voltage excitation. Data were collected and analyzed using CorrView, Corr-Ware, Zplot and ZView software (Scribner Associates Inc., Cambridge, UK).

Int. J. Mol. Sci. 2014, 15

6986

4. Conclusions 1. 2.

3.

4.

5.

Rosin quaternary ammonium salt derivative LPMQA was synthesized by three steps in this work. The structure of target products was characterized by IR, 1H NMR. The data indicated that the LPMQA cationic surfactant was chemically adsorbed at the magnetite surfaces. The charge-stabilized mechanism is most favorable to explain the interaction between magnetite surface and LPMQA cationic surfactants. The formation of monodisperse nanoparticles with an average size of 12.5 nm confirms that the proposed scheme is an effective method for the preparation of monodisperse magnetic nanoparticles. The Fe3O4 NPs size is in good agreement with the average diameter measured from TEM (12.5 nm). Magnetic nanoparticles were an effective inhibitor for corrosion of steel in 1 M HCl solution and the inhibition efficiency increased with increasing concentration. The inhibition efficiencies, calculated from potentiodynamic polarization curve and EIS measurements are in reasonably good agreement. Magnetic nanoparticles inhibit the corrosion of steel by controlling both anodic and cathodic reactions in 1 M HCl solution and acts as a mixed type inhibitor.

Acknowledgments The authors extend their appreciation to the Deanship of Scientific Research at King Saud University for funding this work through research group no RGP-VPP 148. Author Contributions A.M.A. suggested the idea and contributed the discussion, G.A.E.-M. discussed the corrosion data, A.A.-L. contributed in discussion and S.A.A.-H. performed the experimental work. Conflicts of Interest The authors declare no conflict of interest. References 1. 2.

3.

4.

Maiti, S.; Ray, S.S.; Kundu, A.K. Rosin: A renewable resource for polymers and polymer chemicals. Prog. Polym. Sci. 1989, 14, 297–338. Wang, J.F.; Lin, M.T.; Wang, C.P.; Chu, F.X. Study on the synthesis, characterization, and kinetic of bulk polymerization of disproportionated rosin (β-acryloxyl ethyl) ester. J. Appl. Polym. Sci. 2009, 113, 3757–3765. Wilbon, P.A.; Zheng, Y.J.; Yao, K.J.; Tang, C.B.S. Renewable rosin acid-degradable caprolactone block copolymers by atom transfer radical polymerization and ring-opening polymerization. Macromolecules 2010, 43, 8747–8754. Zheng, Y.J.; Yao, K.J.; Lee, J.; Chandler, D.; Wang, J.F.; Wang, C.P.; Chu, F.X.; Tang, C.B. Well-defined renewable polymers derived from gum rosin. Macromolecules 2010, 43, 5922–5924.

Int. J. Mol. Sci. 2014, 15 5. 6. 7. 8. 9. 10. 11. 12.

13.

14. 15. 16. 17.

18.

19. 20. 21. 22. 23.

6987

Kubo, I.; Muroi, H.; Kubo, A. Structural functions of antimicrobial long-chain alcohols and phenols. Bioorg. Med. Chem. Lett. 2010, 3, 873–880. Menger, F.M.; Littau, C.A. Synthesis and properties of dissymmetric gemini surfactants. J. Am. Chem. Soc. 2010, 113, 1451–1452. Wang, J.F.; Chen, Y.P.; Yao, K.J.; Wilbon, P.A. Robust antimicrobial compounds and polymers derived from natural resin acids. Chem. Commun. 2012, 48, 916–918. Li, J.; Rao, X.P.; Shang, S.B.; Gao, Y.Q.; Song, J. Synthesis and antibacterial activity of amide derivatives from acrylopimaric acid. BioResources 2012, 7, 1961–1971. Atta, A.M.; Elsaeed, A.M. Use of rosin-based nonionic surfactants as petroleum crude oil sludge dispersants. J. Appl. Polym. Sci. 2011, 122, 183–192. Atta, A.M.; El-Mahdy, G.A.; Ismail, H.S.; Al-Lohedan, H.A. Effects of water soluble rosin on the corrosion inhibition of carbon steel. Int. J. Electrochem. Sci. 2012, 7, 11834–11846. El-Mahdy, G.A.; Atta, A.M.; Al-Lohedan, H.A. Water soluble Nonionic Rosin Surfactants As Corrosion Inhibitor of carbon steel in 1M HCl. Int. J. Electrochem. Sci. 2013, 8, 5052–5066. Atta, A.M.; El-Mahdy, G.A.; Al-Azhary, A.A.; Al-Lohedan, H.A. Experimental investigation and theoretical approach on water soluble rosin as corrosion inhibitors. Int. J. Electrochem. Sci. 2013, 8, 1295–1307. Atta, A.M.; Ramadan, A.M.; El-Shafay, K.A.; Mohamed, A.M.; Ragab, N.S.; Fekry, M. Synthesis and properties of nonionic surfactants from rosin-imides maleic anhydride adduct. J. Dispers. Sci. Technol. 2009, 30, 1100–1110. Atta, A.M.; Abdel-Rauf, M.E.; Maysour, N.E.; Gafer, A.K. Water-based oil spill dispersants based on rosin formaldehyde resins. J. Dispers. Sci. Technol. 2010, 31, 583–595. Atta, A.M.; Mansour, R.; Abdou, M.I.; Sayed, A.M. Epoxy resins from rosin acids: Synthesis and characterization. Polym. Adv. Technol. 2004, 15, 514–518. Atta, A.M.; El-Saeed, S.M.; Farag, R.K. New vinyl ester resins based on rosin for coating applications. React. Funct. Polym. 2006, 66, 1596–1608. Atta, A.M.; Elsaeed, A.M.; Farag, R.K.; El-Saeed, S.M. Crosslinked reactive macromonomers based on polyisobutylene and octadecyl acrylate copolymers as crude oil sorber. React. Funct. Polym. 2007, 67, 931–943. Atta, A.M.; El-Mahdy, G.A.; Dyab, A.K.F.; Allohedan, H.A. Application of highly surface active cationic surfactants based on rosin as corrosion inhibitor for tubing steel during acidization of petroleum oil and gas wells. Int. J. Electrochem. Sci. 2013, 8, 9629–9643. Nalwa, H.S. Handbook of Nanostructured Materials and Nanotechnology; Academic Press: San Diego, CA, USA, 2000; Volume 1, p. 75. Voevodin, N.; Balbyshev, V.N.; Khobaib, M.; Donley, M.S. Nanostructured coatings approach for corrosion protection. Prog. Org. Coat. 2003, 47, 416–423. Saji, V.S.; Thomas, J. Nano-materials for corrosion control. Curr. Sci. 2007, 92, 51–55. Abdel Hameed, R.S.; Abu-Nawwasb, A.H.; Shehataa, H.A. Nano-composite as corrosion inhibitors for steel alloys in different corrosive media. Adv. Appl. Sci. Res. 2013, 4, 126–129. Benda, P.; Kalendová, A. Anticorrosion properties of pigments based on ferrite coated zinc particles. Phys. Procedia 2013, 44, 185–194.

Int. J. Mol. Sci. 2014, 15

6988

24. Boinovicha, L.B.; Gnedenkov, S.V.; Alpysbaeva, D.A.; Egorkin, V.S.; Emelyanenko, A.M.; Sinebryukhov, S.L.; Zaretskaya, A.K. Corrosion resistance of composite coatings on low-carbon steel containing hydrophobic and superhydrophobic layers in combination with oxide sublayers. Corros. Sci. 2012, 55, 238–245. 25. Zalich, M.A.; Vadala, M.L.; Riffle, J.S.; Saunders, M.; Pierre, T.G.S. Structural and magnetic properties of cobalt nanoparticles encased in siliceous shells. Chem. Mater. 2007, 19, 6597–6604. 26. Pascal, C.; Pascal, J.L.; Favier, F.; Moubtassim, M.L.E.; Payen, C. Electrochemical synthesis for the control of γ-Fe2O3 nanoparticle size. Morphology, microstructure, and magnetic behavior. Chem. Mater. 1999, 11, 141–147. 27. Lee, H.; Kim, D.K.; Jang, N.K.; Jeong, Y.Y.; Jon, S. Antibiofouling polymer-coated superparamagnetic iron oxide nanoparticles as potential magnetic resonance contrast agents for in vivo cancer imaging. J. Am. Chem. Soc. 2006, 128, 7383–7389. 28. Alexiou, C.; Schmid, R.J.; Jurgons, R. Targeting cancer cells: Magnetic nanoparticles as drug carriers. Eur. Biophys. J. 2006, 35, 446–450. 29. Dresco, P.A.; Zaitsev, V.S.; Gambino, R.J.; Chu, B. Preparation and properties of magnetite and polymer magnetite nanoparticles. Langmuir 1999, 15, 1945–1951. 30. Stokinger, H.E. A review of world literature finds iron oxides noncarcinogenic. Am. Ind. Hyg. Assoc. J. 1984, 45, 127–133. 31. Wu, K.T.; Yao, Y.D.; Wang, C.R.C.; Chen, P.F.; Yeh, E.T. Magnetic Field induced optical transmission study in an iron nanoparticle ferrofluid. J. Appl. Phys. 1999, 85, 5959–5961. 32. Zhang, D.; Tong, Z.; Li, S.; Zhang, X.; Ying, A. Characterization of hollow Fe3O4 nanospheres in a microemulsion. Mater. Lett. 2008, 62, 4053–4055. 33. Jun, Y.W.; Huh, Y.M.; Choi, J.S. Nanoscale size effect of magnetic nanocrystals and their utilization for cancer diagnosis via magnetic resonance imaging. J. Am. Chem. Soc. 2005, 127, 5732–5733. 34. Nomura, A.; Shin, S.; Mehdi, O.O.; Kauffmann, J.M. Preparation, characterization, and application of an enzyme-immobilized magnetic microreactor for flow injection analysis. Anal. Chem. 2004, 76, 5498–5502. 35. Li, F.; Vipulanandan, C.; Mohanty, K.K. Microemulsion and solution approaches to nanoparticle iron production for degradation of trichloroethylene. Colloids. Surf. A 2003, 223, 103–112. 36. Yang, C.; Xing, J.; Guan, Y.; Liu, J.; Liu, H. Synthesis and characterization of superparamagnetic iron nanocomposites by hydrazine reduction. J. Alloys Compd. 2004, 385, 283–287. 37. Zhang, W.-X. Nanoscale iron particles for environmental remediation: An overview. J. Nanopart. Res. 2003, 5, 323–332. 38. Atta, A.M.; Dyab, A.K.F. Coated Magnetite Nanoparticles, Method for the Preparation Thereof and Their Use. EP Patent Application, No.: EP13167616.5, 18 March 2013. 39. Khalil, M.I. Process for Preparing Magnetic (Fe3O4) and Derivatives Thereof. EP 2505558 A1, 16 January 2013. 40. Farajia, M.; Yaminia, Y.; Tahmasebia, E.; Saleha, A.; Nourmohammadianb, F. Cetyltrimethylammonium bromide-coated magnetite nanoparticles as highly efficient adsorbent for rapid removal of reactive dyes from the textile companies’ wastewaters. J. Iran. Chem. Soc. 2010, 7, S130–S144.

Int. J. Mol. Sci. 2014, 15

6989

41. Lan, Q.; Yang, F.; Zhang, S.; Liu, S.; Xu, J.; Sun, D. Synergistic effect of silica nanoparticle and cetyltrimethyl ammonium bromide on the stabilization of O/W emulsions. Colloids Surf. A 2007, 302, 126–135. 42. Cornell, R.M.; Schwertmann, U. The Iron Oxides, 2nd ed.; VCH Publishers: Weinheim, Germany, 2003. 43. McClune, W.F. Powder Diffraction File Alphabetical Index Inorganic Phase; JCPDS: Swarthmore, PA, USA, 1980. 44. Cao, C. On electrochemical techniques for interface inhibitor research. Corros. Sci. 1996, 38, 2073–2082. 45. Qu, Q.; Hao, Z.; Li, L.; Bai, W.; Ding, Z. Synthesis and evaluation of Tris-hydroxymethyl-(2-hydroxybenzylidenamino)-methane as a corrosion inhibitor for cold rolled steel in hydrochloric acid. Corros. Sci. 2009, 51, 569–574. 46. Bentiss, F.; Lebrini, M.; Lagrenée, M. Thermodynamic characterization of metal dissolution and inhibitor adsorption processes in mild steel/2,5-bis(n-thienyl)-1,3,4-thiadiazoles/hydrochloric acid system. Corros. Sci. 2005, 47, 2915–2931. 47. Qu, Q.; Jiang, S.; Bai, W.; Li, L. Effect of ethylenediamine tetraacetic acid disodium on the corrosion of cold rolled steel in the presence of benzotriazole in hydrochloric acid. Electrochim. Acta 2007, 52, 6811–6820. 48. Rames, S.; Rajeswari, S. Corrosion inhibition of mild steel in neutral aqueous solution by new triazole derivatives. Electrochim. Acta 2004, 49, 811–820. 49. Chaieb, E.; Bouyanzer, A.; Hammouti, B.; Benkaddour, M. Inhibition of the corrosion of steel in 1 M HCl by eugenol derivatives. Appl. Surf. Sci. 2005, 246, 199–206. 50. Dadgarnezhad, A.; Sheikhshoaie, I.; Baghaei, F. Corrosion inhibitory effects of a new synthetic symmetrical Schiff-base on carbon steel in acid media. Anti-Corros. Meth. Mater. 2004, 51, 266–271. 51. Shi, X.; Nguyen, T.A.; Suo, Z.; Liu, Y.; Avci, R. Effect of nanoparticles on the anticorrosion and mechanical properties of epoxy coating. Surf. Coat. Technol. 2009, 204, 237–245. 52. Yang, L.H.; Liu, F.C.; Han, E.H. Effects of P/B on the properties of anticorrosive coatings with different particle size. Prog. Org. Coat. 2005, 53, 91–98. 53. Lamaka, S.V.; Zheludkevich, M.L.; Yasakau, K.A.; Serra, R.; Poznyak, S.K.; Ferreira, M.G.S. Nanoporous titania interlayer as reservoir of corrosion inhibitors for coatings with self-healing ability. Prog. Org. Coat. 2007, 58, 127–135. © 2014 by the authors; licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution license (http://creativecommons.org/licenses/by/3.0/).

Copyright of International Journal of Molecular Sciences is the property of MDPI Publishing and its content may not be copied or emailed to multiple sites or posted to a listserv without the copyright holder's express written permission. However, users may print, download, or email articles for individual use.

Synthesis of environmentally friendly highly dispersed magnetite nanoparticles based on rosin cationic surfactants as thin film coatings of steel.

This work presents a new method to prepare monodisperse magnetite nanoparticles capping with new cationic surfactants based on rosin. Core/shell type ...
741KB Sizes 0 Downloads 3 Views